Subscriber access provided by UNIV OF LOUISIANA
Environmental Processes
Underappreciated and complex role of nitrous acid in aromatic nitration under mild environmental conditions: the case of activated methoxyphenols Ana Krofli#, Matej Huš, Miha Grilc, and Irena Grgi# Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.8b01903 • Publication Date (Web): 02 Nov 2018 Downloaded from http://pubs.acs.org on November 5, 2018
Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.
is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.
Page 1 of 34
Environmental Science & Technology
1
Underappreciated and complex role of nitrous acid
2
in aromatic nitration under mild environmental
3
conditions: the case of activated methoxyphenols
4
Ana Kroflič,*,a Matej Huš,*,b,c Miha Grilc,c,d and Irena Grgića
5 6
a
7
Ljubljana, Slovenia
8
b
9
Gothenburg, Sweden
Department of Analytical Chemistry, National Institute of Chemistry, Hajdrihova 19, SI-1000
Department of Physics, Chalmers University of Technology, Fysikgränd 3, SE-412 96
10
c
11
Hajdrihova 19, SI-1000 Ljubljana, Slovenia
12
d
13
Germany
Department of Catalysis and Chemical Reaction Engineering, National Institute of Chemistry,
Institute of Chemical Technology, Leipzig University, Linnéstraße 3, DE-04103 Leipzig,
14
1 ACS Paragon Plus Environment
Environmental Science & Technology
Page 2 of 34
15
ABSTRACT
16
Many ambiguities surround the possible mechanisms of colored and toxic nitrophenols formation
17
in natural systems. Nitration of a biologically and environmentally relevant aromatic compound,
18
guaiacol (2-methoxyphenol), under mild aqueous-phase conditions (ambient temperatures, pH 4.5)
19
was investigated by a temperature-dependent experimental modeling coupled to extensive ab initio
20
calculations to obtain the activation energies of the modeled reaction pathways. The importance of
21
dark non-radical reactions is emphasized, involving nitrous (HNO2) and peroxynitrous (HOONO)
22
acids. Oxidation by HOONO is shown to proceed via a non-radical pathway, possibly involving
23
the
24
MP2/6-31++g(d,p) level, NO2• is shown capable of abstracting a hydrogen atom from the phenolic
25
group on the aromatic ring. In a protic solvent, the corresponding aryl radical can combine with
26
HNO2 to yield OH• and, after a subsequent oxidation step, nitrated aromatic products. The
27
demonstrated chemistry is especially important for understanding the aging of nighttime
28
atmospheric deliquesced aerosol. The relevance should be further investigated in the atmospheric
29
gaseous phase. The results of this study have direct implications for accurate modeling of the
30
burden of toxic nitroaromatic pollutants, and the formation of atmospheric brown carbon and its
31
associated influence on Earth’s albedo and climate forcing.
nitronium
ion
(NO2+)
formation.
Using
quantum
chemical
calculations
at
the
32
2 ACS Paragon Plus Environment
Page 3 of 34
Environmental Science & Technology
33
INTRODUCTION
34
There has been much controversy regarding nitration mechanisms of aromatic compounds,
35
especially in complex and diverse biological and environmental systems. Because of extensive
36
industrial applications, aromatic nitration by electrophilic agents such as the nitronium and
37
nitrosonium ions (NO2+ and NO+, respectively) has in general been recognized as the most
38
thoroughly studied reaction in organic chemistry. However, it usually requires extreme
39
conditions.1 Nevertheless, a mounting debate on the exact mechanism of electrophilic aromatic
40
substitution (SEAr) persists.2-6 Considering natural compartments with a plethora of nitrogen-
41
containing reactive species (NRS; e.g. NO2+, NO+, NOx (NO• and NO2•) and NO3• radicals,
42
nitrous and peroxynitrous acids (HNO2 and HOONO), N2O3, N2O4), nitration pathways of
43
aromatic compounds at mild conditions are mechanistically even less understood.7-12
44
Bedini et al.13 have recently employed density functional theory (DFT) at the B3LYP level for
45
studying light-induced nitration of phenol and 4-chlorophenol in environmental waters. Based
46
on the calculations performed in the gaseous phase without accounting for solvation effects, the
47
authors ruled out the OH•-Ar adduct formation before the addition of NO2• to the aromatic ring.
48
Instead, nitration by NO2• alone was proposed. A direct NO2• addition to the aromatic ring and
49
the formation of the nitration product through the H-atom abstraction in a redox reaction that
50
involves NO2• as an oxidant did not describe the experimental observations appropriately. The
51
formation of a phenoxy radical by the first NO2• molecule and a subsequent recombination with
52
the second NO2• molecule seemed most likely. Several other scientists have also proposed an
53
aromatic nitration through the formation of a phenoxy radical in the atmosphere and biological
54
systems rather than the OH•-Ar adduct formation in the first reaction step of aromatic nitration
55
(note: daytime OH• or nighttime NO3• was always required in the first step of the phenoxy
56
radical formation).14-18 In contrast, Zhang et al.19 suggest a water-assisted addition of NO2• to 3 ACS Paragon Plus Environment
Environmental Science & Technology
Page 4 of 34
57
the OH•-Ar or NO3•-Ar adducts in the gas-phase formation of nitrated polycyclic aromatic
58
hydrocarbons. The reaction between the light-excited nitrophenols and NO2• in the presence of
59
oxygen has further been proposed for the formation of dinitrophenol in atmospheric waters.20
60
In phenol-containing HNO2 solutions, nitro- and nitrosophenols form in the dark even at
61
moderate pH, which is consistent with the formation of intermediates in a direct reaction
62
between phenol and HNO2.21 The same has been observed for catechol (CAT, 1,2-
63
dihydroxybenzene).22 Notwithstanding, we have recently ruled out a direct HNO2-driven
64
mechanism of guaiacol (GUA, 2-methoxyphenol) nitration under similar reaction conditions
65
and proposed that electrophilic nitration pathways prevailed in the dark.10 An analogous
66
mechanism has been suggested for the nitration of phenol by HNO2 in the presence of hydrogen
67
peroxide by Vione et al.23
68
Nitroaromatic compounds are well known for their environmental toxicity, carcinogenicity
69
and mutagenicity, and secondary production in the environment.24,25 It has been shown that
70
nitrophenols might be co-responsible for the decline of remote forests downwind from emission
71
sources.26,27 In addition, atmospheric nitration of aromatic pollutants influences the formation of
72
secondary organic aerosol (SOA), which is known to contribute substantially to the existing gap
73
between field measurements and atmospheric models.28 Being an important contributor to
74
atmospheric brown carbon, nitroaromatics alter the absorption properties of the troposphere and
75
thus affect Earth’s energy balance, and directly contribute to climate forcing.29 Therefore, the
76
formation mechanisms of nitroaromatic pollutants under environmentally-relevant conditions
77
are of the utmost interest.
78
For further mechanistic insight into the processes of interest, extensive quantum chemical
79
calculations were carried out in addition to long-term temperature dependent kinetic modeling.
80
Namely, the classical schemes put forward by kinetic studies mostly account for stable
81
quantifiable intermediates, thus inevitably lumping together several elementary reaction steps in 4 ACS Paragon Plus Environment
Page 5 of 34
Environmental Science & Technology
82
an apparent one. In this work, quantum chemical calculations at the MP2/6-31++g(d,p) level
83
were performed to better understand the mechanisms of electrophilic and radical nitration and
84
nitrosation of aromatic compounds in an aqueous medium. The theoretically calculated
85
activation energies of the rate-determining steps (i.e. the reactions between NRS and aromatic
86
molecules) were used in the recently developed predictive model of GUA nitration in
87
atmospheric waters10 to account for the temperature variation of the experimental data. It is
88
shown that in doing so we describe the new experimental dataset very well, confirming and
89
upgrading the theoretically postulated reaction scheme. We not only expand the current
90
understanding of aqueous-phase aromatic nitration under mild environmental conditions but
91
also propose a new non-radical pathway of phenols nitration in the dark where HNO2 plays a
92
pivotal role.
93
METHODS
94
Computational details. Theoretical quantum chemical calculations were performed using
95
Gaussian 09 program suite30 at the MP2/6-31++g(d,p) level.31-35 Some provisional calculations
96
were also performed with B3LYP density functional theory to check whether our choice of the
97
computational method was warranted. With the B3LYP calculations, we were able to locate
98
most, but not all, of the minima with matching geometries and relative energies to those from
99
the MP2 approach. B3LYP is, however, known to miss some energy minima in organic
100
reactions as it underestimates dispersion interactions. Specifically for aromatic nitration, it has
101
been previously pointed out that B3LYP falls short of identifying all stationary points.3
102
Computationally expensive methods, such as single and double excitation coupled cluster
103
(CCSD), reveal all stationary points and arguably give better thermochemistry and geometries,
104
but are prohibitively time-consuming on larger systems with many stationary points. As a
105
compromise, MP2/6-31++g(d,p) was chosen in our case. A comparison of our results with other 5 ACS Paragon Plus Environment
Environmental Science & Technology
Page 6 of 34
106
CCSD calculations on benzene nitration shows that we have located analogous stationary points
107
and recovered comparable energetics, proving that the choice of the method was reasonable.2
108
As shown by Jurečka et al.,36 the difference in energy between MP2 and CCSD(T) is less than 3
109
kJ mol–1, which we take as a 95% confidence interval in our predictive model for the values
110
obtained from the quantum calculations.
111
To account for the aqueous environment of the reactions under consideration, we used a
112
polarizable continuum model (PCM) with the effective dielectric constant 78.5 (the default
113
parameter for water), where solute cavities are embedded as a set of overlapping spheres.
114
Thermochemistry was calculated with zero-point energy taken into account. We report the
115
energetics and thermodynamic quantities relative to the infinitely separated reactants (GUA
116
derivatives and relevant ions or radicals).
117
The intermediates and transition states (TS) were fully optimized without constraints. TS were
118
obtained with the synchronous transit-guided quasi-Newton method (STQN). Vibrational
119
analysis was used to confirm that intermediates had only real frequencies and were thus located
120
in the minima of the potential energy surface (PES), while TS were saddle points and had
121
exactly one imaginary frequency corresponding to the desired reaction pathway. They were
122
confirmed by integrating the intrinsic reaction coordinate (IRC) in both directions from the TS,
123
ending up in products and reactants, respectively.
124
Recombination of radicals is a non-activated process without a saddle point on the PES. The
125
opposite holds true for fragmentation. However, an activation barrier can be found if one
126
considers that this is an example of a spin-forbidden reaction occurring on two potential energy
127
surfaces. The initial triplet state of two separated radicals “hops” into the singlet state at the
128
minimum energy crossing point (MECP), which is an adiabatic transition state. The reaction
129
barrier was defined as the energy at the MECP relative to the initial state (i.e. separated radicals
130
for recombination or stable molecule for fragmentation).37 6 ACS Paragon Plus Environment
Page 7 of 34
Environmental Science & Technology
131
Laboratory experiments. The temperature dependence of GUA nitration in a moderately
132
acidic NaNO2/H2SO4 solution (pH 4.5) was investigated in the dark and under simulated
133
sunlight conditions. The experiments were performed at four different temperatures (283, 293,
134
298, and 303 K; the data at 298 K are taken from our previous work27) and the proposed
135
reaction model was fitted to the experimental data points with high precision (see also Kinetic
136
modeling section). For details see SI, page S1.
137
Kinetic modeling. The influence of temperature on the kinetics of nitration and nitrosation of
138
GUA and its primary reaction products (i.e. nitro and nitrosoguaiacols) was quantitatively
139
explored by the recently developed kinetic model for the reactions performed at 298 K.10 In
140
order to minimize the correlation between the activation energy and the pre-exponential factor
141
during the regression analysis, kinetic rate constants were determined according to a modified
142
Arrhenius equation:
143
∙ 𝑒𝑥𝑝 ― 𝑘𝑇𝑖 = 𝑘298K 𝑖
144
where k and Ea are the kinetic rate constant and activation energy, respectively, R is the gas
145
constant, and i and T denote the reaction number (see column i in Table 1) and the experimental
146
temperature, respectively. The rate constants for all reactions i at 298 K (𝑘298K ) were taken from 𝑖
147
our previous work and were not subjected to further regression analysis.10 The activation
148
energies accounting for the temperature dependence in the model were determined either by
149
regression analysis (mainly electrophilic nitrosation), by ab initio calculations (regiospecific
150
electrophilic and radical nitration and nitrosation of GUA and its primary reaction products) or
151
both (NO•, NO2•, NO+, and NO2+ formation and termination reactions). The set of differential
152
molar balances postulated according to the proposed reaction scheme was numerically solved in
153
Matlab 7.12.0 (MathWorks, Natick, MA, USA). The Levenberg–Marquardt and Jacobian
(
𝐸a𝑖 1 𝑅 𝑇
(
1
― 298K
))
(1)
7 ACS Paragon Plus Environment
Environmental Science & Technology
Page 8 of 34
154
matrix computation methods were used for the regression analysis (parameters optimization)
155
and the subsequent determination of the 95% confidence intervals.
156
RESULTS
157
Experimental-modeling study.
158
The proposed macroscopic reaction scheme of GUA nitration in a moderately acidic NaNO2
159
solution (above the pKa of nitrous acid) is shown in Figure 1.10,27 So far, the model has been
160
verified with data measured at 298 K, whereas this study represents an extension to the
161
temperature dependence in the range from 283 to 303 K. The experimental data at 283, 293,
162
298, and 303 K together with the modeling results are gathered in Figure 2 (nighttime and
163
daylight conditions), showing a remarkably good agreement and offering convincing evidence
164
that the postulated mechanism and the derived activation barriers are sound. We shortly point
165
out a few observations that led us to extend the study to theoretical computations and
166
importantly contribute to the understanding of the nitration of phenolic compounds in natural
167
systems.
168 169
Figure 1. The proposed macroscopic reaction scheme of guaiacol (GUA) aging in a moderately
170
acidic NaNO2 solution. The reaction pathways according to the electrophilic and radical
171
reaction mechanisms are presented with dashed and solid arrows, respectively. The numbers 8 ACS Paragon Plus Environment
Page 9 of 34
Environmental Science & Technology
172
correspond to the kinetic rate constants gathered in Table 1. The identified reaction products are
173
shown in black: 4-nitroguaiacol (4NG), 6-nitroguaiacol (6NG), and 4,6-dinitroguaiacol (DNG).
174
The nitrosoguaiacol (NOG) and nitronitrosoguaiacol (NONG) sideproducts according to the
175
model are colored grey. (Data used from Ref.27)
176
The activation energies, determining the temperature dependence of the kinetic rate constants
177
at 298 K, were first optimized by the regression analysis. We found that the system was
178
insensitive to the variation of Ea of the radical reactions (1–4 and 13 in Table 1) as long as these
179
activation energies were the same or very similar and the ratio between the respective kinetic
180
rate constants remained unchanged. The activation energy of the NO2• formation is the rate-
181
limiting step and determines the influence of temperature on the global nitration rates, whereas
182
the selectivity of the respective radical reactions is governed by the pre-exponential Arrhenius
183
factors, describing the probability of successful collisions. This is consistent with our past
184
observation that the ratio between the radical nitration rate constants is important, not the
185
absolute values.10 Therefore, the activation energies of the reactions between the aromatic
186
components and each reactive species considered in the model were ultimately determined by
187
quantum chemical calculations and used in the modeling. The activation energies obtained from
188
either the quantum chemical calculations and/or experimental modeling (the latter mainly for
189
lumped reactions) are gathered in Table 1 and match the experimental data very well (Figure 2).
190
In the absence of hydrogen peroxide and illumination, the activation energy of the NO2+
191
formation is the highest among all the considered reactive species (Ea22 = 34 kJ mol–1). This
192
means that the cumulative formation of the electrophilic nitronium ion in the dark is most
193
influenced by the reaction temperature, which results in a limited contribution of this reaction
194
pathway at 283 K. Moreover, a relatively large confidence interval is reported for the rate
195
constant of this lumped reaction (k22, Table S1), implying that the overall influence of the
196
electrophilic aromatic nitration on the measured concentration profiles is weak. An additional 9 ACS Paragon Plus Environment
Environmental Science & Technology
Page 10 of 34
197
consideration from the quantum chemical calculations suggests that the present mechanistic
198
representation at the given experimental conditions is not necessarily exact; a novel mechanism
199
is proposed in the following sections.
200
The addition of H2O2 significantly increases the formation of NO2+ at higher temperatures, but
201
still has a negligible effect at 283 K, which is reflected in the high activation energy (Ea29 = 94
202
kJ mol–1) of the lumped reaction of the conversion of HNO2 and H2O2 into HOONO (and
203
presumably further to NO2+). All lumped reactions of the NRS formation/degradation
204
considered in the model are summarized in Scheme S1 and Table S1.
205
In contrast, the low activation energies for the formation of radicals in the dark (Ea20 = 26 kJ
206
mol–1) and under illumination (Ea32 = 22 kJ mol–1) imply that temperature has relatively little
207
effect on these lumped reactions. The modeled values are mostly consistent with our
208
calculations and the literature data for environmental systems.38 The addition of H2O2 also
209
shows an insignificant effect on the formation of radicals; k27 for the NO2• formation is
210
negligible,10 alongside the high energy barrier required for the formation of NO• (Ea28 = 140 kJ
211
mol–1). This opposes many interpretations of aromatic nitration induced by biologically
212
important HOONO,39 but is, however, consistent with a recent understanding, proposing the
213
non-radical nitration mechanism.12
10 ACS Paragon Plus Environment
Page 11 of 34
Environmental Science & Technology
a)
b)
10 °C
0.08 0.04
0.04 0.00 0.08
20 °C
0.04 0.00 0.08
Conc. (mM)
Conc. (mM)
0.00 0.08
25 °C
0.00 0.08
0.04 0.00 0.08
25 °C
0.00 0.08
30 °C
30 °C
0.04
0.04
214
20 °C
0.04
0.04
0.00
GUA 4NG 6NG DNG
10 °C
0.08
0
10
20
30
40
0.00
0
10
20
30
40
Time (h)
Time (h)
215
Figure 2. Experimental data (symbols) of guaiacol (GUA) nitration in a slightly acidic H2SO4
216
solution (pH 4.5) upon the addition of 1 mM NaNO2 and H2O2 a) in the dark and b) under
217
simulated solar irradiation at 283, 293, 298, and 303 K, and the modeled curves (solid lines)
218
according to the proposed reaction scheme. The following main reaction products were
219
quantified: 4-nitroguaiacol (4NG), 6-nitroguaiacol (6NG), and 4,6-dinitroguaiacol (DNG).
220
(Data at 298 K used from Ref.27)
11 ACS Paragon Plus Environment
Environmental Science & Technology
Page 12 of 34
221
Table 1: The best-fit kinetic rate constants (k, ' denotes the electrophilic reaction mechanism) and
222
activation energies (Ea) reported with a 95% confidence interval valid at the experimental conditions at
223
pH 4.5. A method of obtaining Ea is also reported: calculations at the MP2/6-31++g(d,p) level (ab
224
initio) or regression analysis (RA). The rate constants of the corresponding lumped reactions are shown
225
in Table S1. i
NRS
product
ri
kia
Eai (kJ mol–1)b
method
Basic conditions
226 227 228 229 230
1
NO2•
4NG
k1∙[GUA]∙[NO2•]
(4.01 ± 0.04) × 109 L mol–1 s–1
21 ± 3
ab initio
1'
NO2+
4NG
k 1'∙[GUA]∙[NO2+]
(2.52 ± 0.01) × 105 L mol–1 s–1
15 ± 3
ab initio
2
NO2•
6NG
k 2∙[GUA]∙[NO2•]
(5.74 ± 0.04) × 109 L mol–1 s–1
21 ± 3
ab initio
2'
NO2+
6NG
k 2'∙[GUA]∙[NO2+]
(4.07 ± 0.01) × 105 L mol–1 s–1
48 ± 3
ab initio
3
NO2•
DNG
k 3∙[4NG]∙[NO2•]
(7.04 ± 0.08) × 108 L mol–1 s–1
21 ± 3
ab initio
3'
NO2+
DNG
k 3'∙[4NG]∙[NO2+]
(1.42 ± 0.01) × 101 L mol–1 s–1
63 ± 3
ab initio
4
NO2•
DNG
k 4∙[6NG]∙[NO2•]
(1.190 ± 0.009) × 108 L mol–1 s–1
21 ± 3
ab initio
4'
NO2+
DNG
k 4'∙[6NG]∙[NO2+]
(7.01 ± 0.04) × 102 L mol–1 s–1
39 ± 3
ab initio
5
NO2+, NO2−
DNG
k 5∙[GUA]∙[NO2+]∙[NO2−]
(3.03 ± 0.03) × 103 L2 mol–2 s–1
28 ± 8
RA
6
n.a.
unknown
k 6∙[DNG]
(7 ± 1) × 10–6 s–1
21 ± 5
RA
10
NO•
NOG
k 10∙[GUA]∙[NO•]
(6.65 ± 0.05) × 109 L mol–1 s–1
139 ± 3
ab initio
10'
NO+
NOG
k 10'∙[GUA]∙[NO+]
(5.46 ± 0.04) × 102 L mol–1 s–1
57 ± 4
RA
11
NO•
NO4NG
k 11∙[4NG]∙[NO•]
(9.18 ± 0.08) × 108 L mol–1 s–1
139 ± 3
ab initio
12
NO•
NO6NG
k 12∙[6NG]∙[NO•]
(3.86 ± 0.03) × 109 L mol–1 s–1
139 ± 3
ab initio
13
NO2•
NONG
k 13∙[NOG]∙[NO2•]c
(1.095 ± 0.007) × 1010 L mol–1 s–1
21 ± 3
ab initio
13'
NO2+
NONG
k 13'∙[NOG]∙[NO2+]c
(4.09 ± 0.05) × 104 L mol–1 s–1
33–55d
ab initio
a
Data taken from Refs. 10,27
b
The confidence interval for ab initio values is estimated to be 3 kJ mol–1 based on benchmarking MP2 against CCSD(T)
(see Computational details section).36 c
The model does not distinguish between 4- and 6-nitrosoguaiacol. The apparent kinetic rate constant for nitration of both
NOG isomers is reported.
12 ACS Paragon Plus Environment
Page 13 of 34
231 232 233
d
Environmental Science & Technology
Two distinct activation energies were computationally determined for 4NOG and 6NOG nitration, and Ea corresponding to
the apparent kinetic rate constant of their nitration is dependent on the ratio between both isomers; namely, it falls within the reported interval.
13 ACS Paragon Plus Environment
Environmental Science & Technology
Page 14 of 34
234
Quantum chemical calculations.
235
Electrophilic aromatic substitution: reaction mechanism. In the absence of UV light or other
236
sources of radical species, nitration of aromatic compounds proceeds primarily as an
237
electrophilic substitution reaction and, as a rule, requires a strongly acidic medium.1
238
Nevertheless, an electrophilic mechanism of aromatic nitration has also been suggested at
239
milder conditions.10,23 Although SEAr is one of the most well-known and researched reactions in
240
organic chemistry, the mechanistic subtleties are still being debated. Wheland40 proposed the
241
existence of a protonated benzene derivate (σ-complex), which is the key intermediate in most
242
proposed mechanisms.41-46 How it forms, however, differs markedly among the mechanisms.
243
Some propose the formation of different weakly bound intermediates before the formation of
244
the Wheland intermediate, while others favor a single-electron-transfer pathway. Experimental
245
data are inconclusive but theoretical studies seem to favor the latter option.47,48 Although NO2+
246
and NO+ possess similar geometry, electrochemistry, and activity, they are vastly different in
247
their reactivity towards aromatic compounds.2,49 This is a consequence of different reaction
248
pathways, whereas electrophilic aromatic nitrosation displays fewer intermediates and higher
249
saddle-point energies than nitration, which is described in the following sections.50
250
Electrophilic nitration of GUA was first examined. The nitronium ion can approach GUA
251
perpendicularly, forming a weakly bound π-complex 1 (Figure S1). In this structure, common to
252
all nitration pathways for different aromatic sites, the interaction between the delocalized
253
electron system and nitronium ion is weak and the reactants geometries remain unperturbed.
254
This complex is readily formed in a non-activated step, as it is also the case in electrophilic
255
nitration of benzene.51 It quickly converts through transition states 2* and 3* to two different -
256
reactant complexes 4 and 5, respectively, in a non-rate limiting step. In these structures, the
257
nitronium group assumes a bent structure above the ring, halfway between the C1-C2 or C4-C5
258
atoms, respectively. The following step was found to be rate-limiting.52 Activation energies for 14 ACS Paragon Plus Environment
Page 15 of 34
Environmental Science & Technology
259
the conversion of 4 and 5 through 6* and 7* to 8 and 9 were calculated to be +15 and +48 kJ
260
mol−1, respectively, as shown in Figure 3a. The proton abstraction from these Wheland
261
intermediates to yield the end-products, 4NG and 6NG, is known to be a fast, non-rate limiting
262
step and is mediated by the solvent.3 As mentioned before, we deemed these values to be
263
insensitive to temperature in the investigated regime and used them as the activation energies
264
for the reaction rate constants k1’ and k2’, respectively.
265
It should be noted that the formation of 4NG has a much lower energy barrier than 6NG
266
(Figure 3), from which it follows intuitively that 4NG is kinetically favored over 6NG. In fact,
267
this is not consistent with our experimental results (Figure 2a), which agree with the positional
268
selectivity of diffusion-controlled SEAr reactions (o > p >> m in respect to the most activating
269
hydroxyl group on the aromatic ring).53 Nevertheless, the experimental data can be reconciled
270
with the theoretical calculations if the stability of both intermediates is taken into account.
271
Namely, the precursor -complex of 6NG (5) is thermodynamically much more stable than that
272
of 4NG (4). As provisionally shown in Figure 3a, 4 can and does quickly convert into 5 with
273
virtually no activation barrier. This means that although having to traverse a higher saddle point
274
7*, 9 is preferably formed as its precursor 5 exists in larger amounts than 4. Still, at higher
275
temperatures, the preferential formation of the thermodynamically more stable product (4NG) is
276
expected.
277
Electrophilic nitration of the nitrated GUA derivatives also proceeds at the given reaction
278
conditions; both 4NG and 6NG are readily nitrated into DNG (reactions k3’ and k4’). The
279
mechanism includes analogous intermediates and transition states as in the nitration of GUA
280
(for details and structures 10–29 see Figures S2–S3). Activation energies of +63 and
281
+39 kJ mol−1 were calculated for the nitration of 4NG and 6NG, respectively, as shown in
282
Figure 3c. Nitration of 6NG is predicted to be faster than that of 4NG, which is consistent with
15 ACS Paragon Plus Environment
Environmental Science & Technology
Page 16 of 34
283
our experimental results (see Figure 2b and compare the decays of both products after GUA has
284
been completely consumed).
285 286
Figure 3. Graphical representation of the potential energy surface for the electrophilic nitration
287
of
288
(c) 4-nitrosoguaiacol (4NOG) and 6-nitrosoguaiacol (6NOG), and for the (d) electrophilic
289
nitrosation of guaiacol (GUA). Only nitration/nitrosation on ortho (red solid line) and para
290
(blue dashed line) positions are considered. The compounds labeled as in Figures S1–4.
(a) guaiacol
(GUA),
(b) 4-nitroguaiacol
(4NG)
and
6-nitroguaiacol
(6NG)
291
Electrophilic nitration of the nitroso-substituted GUA was investigated in the same manner
292
(for details see SI, pages S5–S6). The formation of 4-nitro-6-nitrosoguaiacol (Ea = +33 kJ
293
mol−1) is again predicted to occur faster than that of 6-nitro-4-nitrosoguaiacol (Ea = +55 kJ
294
mol−1), which is presented in Figure 3b. Since our experimental apparatus was not able to
295
distinguish between the two products, their kinetics was lumped in one apparent reaction k13’.
16 ACS Paragon Plus Environment
Page 17 of 34
Environmental Science & Technology
296
These calculated activation energies were used in our kinetic model, while the activation
297
energy of the lumped reactions involved in NO2+ formation from HNO2 was determined by the
298
regression analysis.
299
Similarly to what has been described for nitration, electrophilic aromatic nitrosation begins
300
with a weakly bound π-complex between GUA and NO+ (30, Figure S4); the nitrosonium ion is
301
positioned above the ring and its nitrogen atom is interacting with the π-electrons. This adduct
302
then transforms into protonated nitrosoguaiacols (33 and 34) in a single elementary rate-
303
determining step (see Figure 3d for the PES). As already pointed out by Skokov and Wheeler,50
304
the transient Wheland structure in the reaction of nitrosation is not a true intermediate but
305
merely a transition state (31* and 32*). The activation energies in this mechanism are
306
+263 kJ mol−1 and +309 kJ mol−1 for the formation of 4NOG and 6NOG, respectively,
307
corresponding to the lumped k10’ in our model. This is consistent with known facts and our
308
experimental findings that electrophilic nitrosation proceeds several orders of magnitudes
309
slower than nitration, and in line with the theoretical results from Gwaltney et al.2 The final
310
abstraction of proton is again a fast reaction.
311
In our model, however, the activation energy of this reaction was determined by regression
312
analysis to fit the experimental results better. Note that the authors had avoided the introduction
313
of OH• in the predictive model on purpose, i.e. to reduce the degrees of freedom and improve
314
the predictive power of the model with regard to aromatic nitration.10 This compensates for the
315
high activation energies of electrophilic nitrosation with the Ea of hydroxylation (21 kJ mol−1)
316
and results in the experimentally determined Ea10’ of 57 kJ mol−1.
317
Homolytic aromatic substitution: the reaction mechanism. Homolytic aromatic substitution
318
reactions prevail under daytime conditions because radical species mainly form photolytically;
319
including OH• as the most important radical in the environment. This chemistry is, however,
320
always accompanied by electrophilic and other non-radical reactions (e.g. oxidation-reduction 17 ACS Paragon Plus Environment
Environmental Science & Technology
Page 18 of 34
321
reactions) that are insensitive to light. There are several possible pathways for the formation and
322
interconversion of radicals. However, radical reactions are generally fast enough for their
323
observed rate to be determined by transport phenomena (see for instance the calculated barriers
324
for the recombination reactions below). With this in mind, an approximate uniform value of 21
325
kJ mol–1 was chosen as the activation energy of radical reactions in the predictive model (Table
326
1), which is consistent with the quantum chemistry calculations (vide infra the reaction of GUA
327
with NO2•). For clarity, only the reactions pertaining to the para site are elaborated henceforth.
328
The ortho site is reactive in a similar fashion.
329
Initiation. In the environment, GUA can react with various radical species, including OH•,
330
NO3• and, theoretically, NO• and NO2•. When OH• is present in large quantities, it readily and
331
barrierlessly attaches to aromatic carbon atoms due to high stabilization energy. The results of
332
ab initio calculations presented in the SI show that there is little preference as to where OH•
333
binds (structures 35–40 in Figure S5). It is slightly more favorable, however, for OH• to bind to
334
the hydroxyl-bearing carbon atom (C1), yielding 35. This agrees with the subsequent findings
335
that radical reactions begin with H• being cleaved off the OH group (vide infra). The
336
stabilization energies of these isomers with respect to the isolated reactants range from −65 to
337
−86 kJ mol−1.
338
In contrast to the general perception that NO2• is too weak to react with aromatic compounds,
339
our calculations support the work of Bedini et al.13 and demonstrate that NO2• is not only
340
capable of abstracting hydrogen atoms from GUA while turning itself into HNO2, but also that
341
this is the prevailing mechanism of the GUA-derived radicals formation in the dark. NO2• most
342
likely abstracts the hydrogen atom from the GUA hydroxyl group. It first forms a weakly bound
343
(3 kJ mol–1) planar adduct (41), which has the activation barrier of 21 kJ mol–1 (42*) for
344
hydrogen abstraction, yielding a phenoxy radical (43, Figures S8–S9) and HNO2. As this
345
reaction has the lowest barrier among all involving radicals and GUA, we believe this is the 18 ACS Paragon Plus Environment
Page 19 of 34
Environmental Science & Technology
346
main pathway in the radical nitration mechanism initiated by NO2•. Thus, this activation barrier
347
was used in our mechanistic model to describe the radical reactions. Interestingly, despite OH•
348
being much more reactive than NO2•, the activation barrier for the phenoxy radical formation
349
following the attack of OH• is much higher (69 kJ mol–1; see SI, Figures S6 and S8, and page S8
350
for details); this can be explained by the high stabilization energy of the GUA-OH• adducts (35–
351
40).
352
Alternatively, NO2• might abstract the aromatic hydrogen atom. The adduct of NO2• at the
353
para position of GUA (44) does not exhibit any stabilization (less than 1 kJ mol–1) and has a
354
prohibitively high activation energy of 150 kJ mol–1 (45*), which yields non-negligible amounts
355
of aryl species via direct hydrogen abstraction (46, Figures S8–S9). A solvent-mediated
356
isomerization from the phenoxy radical is thus a more likely way of the aryl radical formation.
357
It is even less likely for NO2• to substitute the hydroxyl group. A weakly bound π complex
358
(47) with negligible stabilization energy would have to overcome the barrier (48*) of 169 kJ
359
mol–1 for NO2• to displace the hydroxyl group which would migrate to the adjacent carbon atom
360
(49). The ensuing 2-nitroanisol (50) and OH• radical are by 165 kJ mol–1 less stable than GUA
361
and NO2• when completely separated; this is expected as NO2• is a better leaving group than
362
OH•. It is also not possible for NO2• to insert directly to the para position via transition state 51*
363
and yield a hydrogenated 4NG•H (52) as the required activation energy exceeds 220 kJ mol–1
364
(Figure S10).
365
NO3•-driven nitration is one of the generally accepted pathways for the formation of
366
nitroaromatic compounds in the environment. The NO3• radical is much more reactive than
367
NO2•; it is predominantly a nighttime oxidant (it rapidly photolyses during the day) but was not
368
present in our experimental system. For the sake of completeness, we studied its reactions as
369
well (Figures S11–S12). To sum up the results of the quantum chemical calculations presented
370
in the SI (page S12): if NO3• were present in the reaction mixture, it would considerably speed 19 ACS Paragon Plus Environment
Environmental Science & Technology
Page 20 of 34
371
up the reaction, generating 43 and 46 (and the corresponding isomers) orders of magnitudes
372
faster.
373
NO• behaves in a similar fashion as NO2•; however, due to its lower reactivity, the reactions
374
are much slower. For instance, upon weakly interacting with the hydroxyl hydrogen (53), the
375
activation barrier for the abstraction (54*) of this hydrogen is 139 kJ mol–1. Moreover, the
376
ensuing HNO and phenoxy radical (43) are by 126 kJ mol–1 less stable than the reactants. Also,
377
the substitution of the hydroxyl group by NO• (55*) would require surmounting the activation
378
energy of 214 kJ mol–1 to displace the OH group to the adjacent carbon atom (56) and
379
ultimately yield 2-nitrosoanisol (57). Similarly, a direct insertion of NO• in para position (58*)
380
to yield a hydrogenated 4NOG•H (59) is inaccessible with a barrier of 213 kJ mol–1 (Figure
381
S13).
382
Propagation. Once the aryl radical of GUA (such as 46) is formed through the solvent-
383
mediated isomerization of the phenoxy radical (possibly via a carbocation radical intermediate;
384
see Figure S14 and Hemberger et al.54) or by the OH• attack on the unsubstituted carbon atoms
385
(cf. Figures S6–S7), it can react with HNO2 or, less readily, HNO3 (for reactions with radical
386
species vide infra) and form nitroso and nitro products. With HNO2, the hydrogenated 4NG•H
387
(52) is formed extremely fast as the activation barrier (60*) is only 9 kJ mol–1. HNO3, however,
388
has to overcome a barrier of 73 kJ mol−1 (61*) to form the 4NG•OH adduct with the hydroxyl
389
group attached to the adjacent (in position meta) carbon atom (62). For structures, see Figure
390
S15.
391
Lastly, the hydrogenated 4NG•H (52) converts into either 4NG or 4NOG. To form 4NG, the
392
hydrogen atom from the nitro group must first migrate (63*) to the adjacent carbon atom (64) or
393
be cleaved off by another radical. The activation energy for its migration is high (162 kJ mol–1)
394
and it is thus more likely that the hydrogen is cleaved off by NO• (65*) or NO2• (66*) with the
395
activation barriers of 44 and 39 kJ mol−1, respectively. Alternatively, the N-O bond in the 20 ACS Paragon Plus Environment
Page 21 of 34
Environmental Science & Technology
396
protonated nitro group can also break, resulting in the migration of the OH moiety (67*) to the
397
adjacent carbon atom (68), which ultimately yields 4NOG. Such rearrangement has a lower
398
activation barrier (126 kJ mol–1) and helps explain why nitroso products are formed despite the
399
NO• radical being much less reactive than NO2•. Similarly as 4NG•H, the hydrogenated
400
4NOG•H (59) can lose its hydrogen atom through the attack of NO• (69*) or NO2• (70*) with the
401
activation barriers of 89 and 58 kJ mol−1. See Figure S16 for structures. The formed
402
nitrosoaromates are readily oxidized to their respective nitro analogoues.55 This has been
403
recently considered as one of the competitive mechanisms in the CAT nitration under similar
404
experimental conditions.22
405
Termination. Ultimately, radical reactions come to a halt when two radicals recombine.
406
Activation barriers, defined as the energy of the MECP (Figure S17) for the association of aryl
407
radical (46) and the nitrogen-containing radicals (i.e. NO2• and NO•), are gathered in the SI
408
(page S16), together with the corresponding transition states presented in Figure S18. However,
409
the termination reactions (recombination) are so fast that they are essentially transport-limited
410
and the barriers given in the SI are not rate-determining for our model. As expected for radical
411
reactions, the termination steps occur when the concentration of the radicals increases relative to
412
the concentration of the reactants, which typically happens in the latter stages of the reaction.
413
DISCUSSION
414
Pathways of aromatic nitration. NO2• is shown capable of reacting with activated phenol,
415
which has not been generally accepted yet. Still, daytime OH• and nighttime NO3• are much
416
stronger oxidants. NO• is not reactive enough on its own: however, it is important in the
417
subsequent reaction steps. For a complete scheme of the studied radical reactions, refer to
418
Figure 4.
21 ACS Paragon Plus Environment
Environmental Science & Technology
Page 22 of 34
419
To summarize the most important conclusions regarding the studied nitration, the relevance of
420
which for atmospheric nighttime chemistry should be considered in the future; if NO3• were
421
present in the solution, it would be the main source of the GUA-derived radicals in the dark. In
422
its absence, NO2• solely is responsible for producing reactive aryl radicals and does not directly
423
substitute the groups on the aromatic ring. NO2• rather reacts with GUA and forms the
424
corresponding phenoxy radical (GUA-O•), which further isomerizes to its aryl analogue
425
(GUA•-OH) in a solvent-mediated step. It can also recombine with other radical species in
426
solution. Nevertheless, the coupled oxidation-reduction reaction involving HNO2 substantially
427
increases the importance of the nighttime nitration of activated phenols by NO2• in low
428
concentration. GUA•-OH can combine with HNO2 to form 4- and 6-nitrosoguaiacols (NOG),
429
which are relatively easily oxidized to the respective nitroderivatives (4NG and 6NG; e.g. with
430
HNO2 or NO2•, concomitantly forming NO•, which is not strong enough to further react with
431
present aromatics). See Scheme 1 for the set of relevant reactions.
22 ACS Paragon Plus Environment
Page 23 of 34
Environmental Science & Technology
432 433
Figure 4. A complete reaction scheme of the studied radical reactions involving GUA and NRS.
434
For clarity, only the attack on para site and formation of mono-substituted products is
435
considered. The colored values in italics (green for accessible, red for inaccessible) show the
436
activation energies of the steps in kJ mol−1, as obtained from our ab initio calculations.
437
Thermolysis of HNO2 and the consumption of the unstable by-product HNO are also shown.
438
In the presence of H2O2, however, it has recently been shown feasible that an electrophilic
439
nitrogen, presumably NO2+, forms under mild atmospheric conditions. Even upon the in-situ
440
formation of H2O2 from oxygen in a redox system of (hydro)quinones, such as CAT.10 We
441
generally confirm the electrophilic mechanism of GUA nitration in the presence of H2O2.
442
Notwithstanding, as only traces of CAT impurity were present in the GUA solution, the dark
443
reaction performed in the absence of oxygen and H2O2 suggests rather the existence of an 23 ACS Paragon Plus Environment
Environmental Science & Technology
Page 24 of 34
444
equilibrium which opposes the GUA nitration if the solution is not aerated. The corresponding
445
experiment performed under N2 atmosphere (O2 was completely expelled from the reaction
446
medium) is shown in Figure S19 and the mismatch between the experimental data and the
447
predictive model at longer reaction times can be nicely explained by the reactions summarized
448
in Scheme 1.
449
2 HNO2 ↔ NO• + NO2• + H2O
450
GUA + NO2• ↔ GUA-O• + HNO2
451
GUA-O• ↔ GUA•-OH
452
GUA•-OH + HNO2 → NOG + OH•
453
NOG + HNO2 → NG + NO• + H+
454
NO•(aq) + [O] ↔ HNO2
455
Scheme 1. Complex role of HNO2 in guaiacol nitration.
456
It is well known that HNO2 reversibly thermolyses to NO• and NO2• in aqueous solutions,56
457
which we also confirmed computationally (see the SI, page S18). The values agree very well
458
with Ea of the lumped reaction k20 for the formation of NO• and NO2• from HNO2 in the
459
equilibrium with NO2− (Table S1). In aerated aqueous solutions, HNO2 can regenerate from
460
NO• in the reaction with O2, whereas in the absence of oxygen NO• accumulates and stops the
461
production of NO2• by shifting the equilibrium towards HNO2. This agrees with our
462
observations, where in the absence of oxygen, GUA nitration initially proceeds, and stops after
463
some time.
464
A one-step oxidation of GUA by HNO2 (resulting in 43 or 46), however, was neither observed
465
experimentally nor could any suitable transition state be located computationally. The excess
466
concentration of HNO2 applied in this work would not allow such experimental system to 24 ACS Paragon Plus Environment
Page 25 of 34
Environmental Science & Technology
467
respond in a switch-like manner (see Figure 2b), even though GUA is presumably more easily
468
oxidized than its nitrated analogues, which contain electron withdrawing substituents. Instead of
469
the delayed formation of DNG, its exponential production from the beginning of the experiment
470
would be expected, which could slightly accelerate after GUA would have been completely
471
consumed.
472
Environmental relevance. An approach coupling extensive experimental data collection, a
473
powerful experimental-modeling investigation and a comprehensive battery of quantum
474
chemical calculations on the mechanism of electrophilic and radical aromatic reactions
475
involving diverse NRS was employed to provide new insights into the nitration of activated
476
aromatic compounds under environmentally and biologically relevant conditions. The
477
represented results strongly support the current state of understanding that a HOONO induced
478
oxidation proceeds via the non-radical mechanism. Furthermore, a special role of HNO2 in the
479
dark formation of phenoxy radicals (through NO2• formation) and their further transformation
480
into nitrophenols, yielding also OH• (by reaction with HNO2 itself), is emphasized. In short,
481
after the phenoxy radicals are formed by trace NO2•, they isomerize to the aryl radicals that can
482
react with HNO2 to the immediate precursors of nitroso- and nitrophenols. The intermediate
483
nitrosophenols are ultimately oxidized to their corresponding nitrated products (again possibly
484
with HNO2), which are potentially toxic for living organisms and, if present in the atmosphere,
485
absorb solar and terrestrial irradiation. The overall influence of temperature on aromatic
486
nitration is small as the activation energies of the respective reactions are low to moderate.
487
Although the presented chemistry is especially important at low pH (e.g. in deliquesced
488
atmospheric aerosols, pH-dependent data not shown), it is demonstrated to be relevant also at
489
milder conditions, above the pKa of HNO2, which has not been generally recognized yet. The
490
authors want to underline the possible importance of the demonstrated HONO-mediated
491
chemistry also in the atmospheric gaseous phase, which brings novel aspects about the role of 25 ACS Paragon Plus Environment
Environmental Science & Technology
Page 26 of 34
492
this emerging atmospheric pollutant. These findings, if included into atmospheric models,
493
would improve the description of the likely underestimated dark formation of nitrophenols in
494
the atmospheric models.
495 496
ASSOCIATED CONTENT
497
Supporting Information
498
The Supporting Information is available free of charge on the ACS Publications website at DOI:
499
Materials and methods; set of lumped reactions; structures of intermediates and transition
500
states; mechanisms of electrophilic nitration for 4NOG, 6NOG, 4NG and 6NG;
501
mechanisms of radical reactions involving OH•, NO2• and NO3•; activation energy
502
determination for radical reactions; coordinates of the located stationary points; phenoxy
503
radical isomerization, experiment in the absence of oxygen.
504
AUTHOR INFORMATION
505
Corresponding Authors
506
*E-mails:
[email protected] and
[email protected].
507
Notes
508
The authors declare no competing financial interest.
509
ACKNOWLEDGMENTS
510
The authors acknowledge the financial support from the Slovenian Research Agency (research core
511
funding Nos. P1-0034 and P2-0152). M.H. also wishes to thank the Knut and Alice Wallenberg
512
Foundation (Project 2015.0057) for funding and Matthias Vandichel for fruitful discussions regarding
513
the chemistry of radicals. 26 ACS Paragon Plus Environment
Page 27 of 34
Environmental Science & Technology
514
REFERENCES
515
(1) Mortier, J. Arene Chemistry: Reaction Mechanisms and Methods for Aromatic Compounds;
516
Wiley: 2015.
517
(2) Gwaltney, S. R.; Rosokha, S. V.; Head-Gordon, M.; Kochi, J. K. Charge-transfer
518
mechanism for electrophilic aromatic nitration and nitrosation via the convergence of (ab initio)
519
molecular-orbital and Marcus-Hush theories with experiments. J. Am. Chem. Soc. 2003, 125
520
(11), 3273-3283.
521
(3) Esteves, P. M.; Carneiro, J. W. D.; Cardoso, S. P.; Barbosa, A. G. H.; Laali, K. K.; Rasul,
522
G.; Prakash, G. K. S.; Olah, G. A. Unified mechanistic concept of electrophilic aromatic
523
nitration: Convergence of computational results and experimental data. J. Am. Chem. Soc. 2003,
524
125 (16), 4836-4849.
525
(4) Koleva, G.; Galabov, B.; Wu, J. I.; Schaefer, H. F.; Schleyer, P. V. Electrophile affinity: a
526
reactivity measure for aromatic substitution. J. Am. Chem. Soc. 2009, 131 (41), 14722-14727.
527
(5) Galabov, B.; Nalbantova, D.; Schleyer, P. V.; Schaefer, H. F. Electrophilic Aromatic
528
Substitution: New Insights into an Old Class of Reactions. Acc. Chem. Res 2016, 49 (6), 1191-
529
1199.
530
(6) Nieves-Quinones, Y.; Singleton, D. A. Dynamics and the Regiochemistry of Nitration of
531
Toluene. J. Am. Chem. Soc. 2016, 138 (46), 15167-15176.
532
(7) Birkmann, B.; Owens, B. T.; Bandyopadhyay, S.; Wu, G.; Ford, P. C. Synthesis of a nitro
533
complex of Ru-III(salen): Unexpected aromatic ring nitration by a nitrite salt. J. Inorg.
534
Biochem. 2009, 103 (2), 237-242.
535
(8) Elias, G.; Mincher, B. J.; Mezyk, S. P.; Cullen, T. D.; Martin, L. R. Anisole nitration during
536
gamma-irradiation of aqueous nitrite and nitrate solutions: free radical versus ionic mechanisms.
537
Environ. Chem. 2010, 7 (2), 183-189. 27 ACS Paragon Plus Environment
Environmental Science & Technology
Page 28 of 34
538
(9) Vione, D.; Maurino, V.; Minero, C.; Pelizzetti, E. Reactions induced in natural waters by
539
irradiation of nitrate and nitrite ions. In Environmental Photochemistry Part II; Boule, P.,
540
Bahnemann, D. W., Robertson, P. K. J., Eds.; Springer Berlin Heidelberg: 2005; pp 221-253.
541
(10) Kroflič, A.; Grilc, M.; Grgić, I. Unraveling pathways of guaiacol nitration in atmospheric
542
waters: Nitrite, a source of reactive nitronium ion in the atmosphere. Environ. Sci. Technol.
543
2015, 49 (15), 9150-9158.
544
(11) Rodenko, B.; Koch, M.; van der Burg, A. M.; Wanner, M. J.; Koomen, G. J. The
545
mechanism of selective purine C-nitration revealed: NMR studies demonstrate formation and
546
radical rearrangement of an N7-nitramine intermediate. J. Am. Chem. Soc. 2005, 127 (16),
547
5957-5963.
548
(12) Koppenol, W. H.; Bounds, P. L.; Nauser, T.; Kissner, R.; Ruegger, H. Peroxynitrous acid:
549
controversy and consensus surrounding an enigmatic oxidant. Dalton Trans. 2012, 41 (45),
550
13779-13787.
551
(13) Bedini, A.; Maurino, V.; Minero, C.; Vione, D. Theoretical and experimental evidence of
552
the photonitration pathway of phenol and 4-chlorophenol: A mechanistic study of
553
environmental significance. Photochem. Photobiol. Sci. 2012, 11 (2), 418-424.
554
(14) Olariu, R. I.; Klotz, B.; Barnes, I.; Becker, K. H.; Mocanu, R. FT-IR study of the ring-
555
retaining products from the reaction of OH radicals with phenol, o-, m-, and p-cresol. Atmos.
556
Environ. 2002, 36 (22), 3685-3697.
557
(15) Berndt, T.; Böge, O. Gas-phase reaction of OH radicals with phenol. Phys. Chem. Chem.
558
Phys. 2003, 5 (2), 342-350.
559
(16) Xu, C.; Wang, L. M. Atmospheric Oxidation Mechanism of Phenol Initiated by OH
560
Radical. J. Phys. Chem. A 2013, 117 (11), 2358-2364.
28 ACS Paragon Plus Environment
Page 29 of 34
Environmental Science & Technology
561
(17) Papavasileiou, K. D.; Tzima, T. D.; Sanakis, Y.; Melissas, V. S. A DFT study of the nitric
562
oxide and tyrosyl radical interaction: A proposed radical mechanism. Chemphyschem 2007, 8
563
(18), 2595-2602.
564
(18) Platz, J.; Nielsen, O. J.; Wallington, T. J.; Ball, J. C.; Hurley, M. D.; Straccia, A. M.;
565
Schneider, W. F.; Sehested, J. Atmospheric chemistry of the phenoxy radical, C6H5O(center
566
dot): UV spectrum and kinetics of its reaction with NO, NO2, and O-2. J. Phys. Chem. A 1998,
567
102 (41), 7964-7974.
568
(19) Zhang, Q.; Gao, R.; Xu, F.; Zhou, Q.; Jiang, G.; Wang, T.; Chen, J.; Hu, J.; Jiang, W.;
569
Wang, W. Role of Water Molecule in the Gas-Phase Formation Process of Nitrated Polycyclic
570
Aromatic Hydrocarbons in the Atmosphere: A Computational Study. Environ. Sci. Technol.
571
2014, 48 (9), 5051−5057.
572
(20) Vione, D.; Maurino, V.; Minero, C.; Pelizzetti, E. Aqueous atmospheric chemistry:
573
formation of 2,4-dinitrophenol upon nitration of 2-nitrophenol and 4-nitrophenol in solution.
574
Environ. Sci. Technol. 2005, 39 (20), 7921-7931.
575
(21) Vione, D.; Belmondo, S.; Carnino, L. A kinetic study of phenol nitration and nitrosation
576
with nitrous acid in the dark. Environ. Chem. Lett. 2004, 2 (3), 135-139.
577
(22) Vidović, K.; Lašič Jurković, D.; Šala, M.; Kroflič, A.; Grgić, I. Nighttime aqueous phase
578
formation of nitrocatechols (1,2 dihydroxynitrobenzenes) in the atmospheric condensed phase.
579
Environ. Sci. Technol. 2018, 52 (17), 9722-9730.
580
(23) Vione, D.; Maurino, V.; Minero, C.; Borghesi, D.; Lucchiari, M.; Pelizzetti, E. New
581
processes in the environmental chemistry of nitrite. 2. The role of hydrogen peroxide. Environ.
582
Sci. Technol. 2003, 37 (20), 4635-4641.
583
(24) Kovacic, P.; Somanathan, R. Nitroaromatic compounds: environmental toxicity,
584
carcinogenicity, mutagenicity, therapy and mechanism. J. Appl. Toxicol. 2014, 34 (8), 810-824.
29 ACS Paragon Plus Environment
Environmental Science & Technology
Page 30 of 34
585
(25) Pflieger, M.; Kroflic, A. Acute toxicity of emerging atmospheric pollutants from wood
586
lignin due to biomass burning. J. Hazard. Mater. 2017, 338, 132-139.
587
(26) Rippen, G.; Zietz, E.; Frank, R.; Knacker, T.; Klopffer, W. Do airborne nitrophenols
588
contribute to forest decline. Environ. Technol. Lett. 1987, 8 (10), 475-482.
589
(27) Kroflič, A.; Grilc, M.; Grgić, I. Does toxicity of aromatic pollutants increase under remote
590
atmospheric conditions? Sci. Rep. 2015, 5, 8859.
591
(28) Yee, L. D.; Kautzman, K. E.; Loza, C. L.; Schilling, K. A.; Coggon, M. M.; Chhabra, P. S.;
592
Chan, M. N.; Chan, A. W. H.; Hersey, S. P.; Crounse, J. D.; Wennberg, P. O.; Flagan, R. C.;
593
Seinfeld, J. H. Secondary organic aerosol formation from biomass burning intermediates:
594
phenol and methoxyphenols. Atmos. Chem. Phys. 2013, 13 (16), 8019-8043.
595
(29) Lin, P.; Bluvshtein, N.; Rudich, Y.; Nizkorodov, S. A.; Laskin, J.; Laskin, A. Molecular
596
Chemistry of Atmospheric Brown Carbon Inferred from a Nationwide Biomass Burning Event.
597
Environ. Sci. Technol. 2017, 51 (20), 11561-11570.
598
(30) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J.
599
R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G. A.; Nakatsuji, H.; Caricato, M.; Li,
600
X.; Hratchian, H. P.; Izmaylov, A. F.; Bloino, J.; Zheng, G.; Sonnenberg, J. L.; Hada, M.;
601
Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao,
602
O.; Nakai, H.; Vreven, T.; Montgomery, J. A., Jr.; Peralta, J. E.; Ogliaro, F.; Bearpark, M.;
603
Heyd, J. J.; Brothers, E.; Kudin, K. N.; Staroverov, V. N.; Kobayashi, R.; Normand, J.;
604
Raghavachari, K.; Rendell, A.; Burant, J. C.; Iyengar, S. S.; Tomasi, J.; Cossi, M.; Rega, N.;
605
Millam, J. M.; Klene, M.; Knox, J. E.; Cross, J. B.; Bakken, V.; Adamo, C.; Jaramillo, J.;
606
Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski,
607
J. W.; Martin, R. L.; Morokuma, K.; Zakrzewski, V. G.; Voth, G. A.; Salvador, P.; Dannenberg,
608
J. J.; Dapprich, S.; Daniels, A. D.; Farkas, O.; Foresman, J. B.; Ortiz, J. V.; Cioslowski, J.; Fox,
609
D. J. Gaussian 09, Revision E.01; Gaussian, Inc., Wallingford CT, 2009. 30 ACS Paragon Plus Environment
Page 31 of 34
Environmental Science & Technology
610
(31) Head-Gordon, M.; Pople, J. A.; Frisch, M. J. MP2 energy evaluation by direct methods.
611
Chem. Phys. Lett. 1988, 153 (6), 503-506.
612
(32) Saebo, S.; Almlof, J. Avoiding the integral storage bottleneck in LCAO calculations of
613
electron correlation. Chem. Phys. Lett. 1989, 154 (1), 83-89.
614
(33) Frisch, M. J.; Head-Gordon, M.; Pople, J. A. A direct MP2 gradient-method. Chem. Phys.
615
Lett. 1990, 166 (3), 275-280.
616
(34) Frisch, M. J.; Head-Gordon, M.; Pople, J. A. Semidirect algorithms for the MP2 energy and
617
gradient. Chem. Phys. Lett. 1990, 166 (3), 281-289.
618
(35) Head-Gordon, M.; Head-Gordon, T. Analytic MP2 frequencies without 5th-order storage –
619
Theory and application to bifurcated hydrogen-bonds in the water hexamer. Chem. Phys. Lett.
620
1994, 220 (1-2), 122-128.
621
(36) Jurečka, P.; Šponer, J.; Černý, J.; Hobza, P. Benchmark database of accurate (MP2 and
622
CCSD(T) complete basis set limit) interaction energies of small model complexes, DNA base
623
pairs, and amino acid pairs. Phys. Chem. Chem. Phys. 2006, 8 (17), 1985-1993.
624
(37) Harvey, J. N. Understanding the kinetics of spin-forbidden chemical reactions. Phys.
625
Chem. Chem. Phys. 2007, 9 (3), 331-343.
626
(38) McKay, G.; Rosario-Ortiz, F. L. Temperature Dependence of the Photochemical Formation
627
of Hydroxyl Radical from Dissolved Organic Matter. Environ. Sci. Technol. 2015, 49 (7), 4147-
628
4154.
629
(39) Gunaydin, H.; Houk, K. N. Mechanisms of Peroxynitrite-Mediated Nitration of Tyrosine.
630
Chem. Res. Toxicol. 2009, 22 (5), 894-898.
631
(40) Wheland, G. W. A quantum mechanical investigation of the orientation of substituents in
632
aromatic molecules. J. Am. Chem. Soc. 1942, 64 (4), 900-908.
31 ACS Paragon Plus Environment
Environmental Science & Technology
Page 32 of 34
633
(41) Hughes, E. D.; Ingold, C. K.; Reed, R. I. Kinetics and mechanism of aromatic nitration.
634
Part II. Nitration by the nitronium ion, NO2+, derived from nitric acid. J. Chem. Soc. 1950,
635
2400-2440
636
(42) Olah, G. A.; Kuhn, S. J.; Flood, S. H. Aromatic substitution. VIII.1 Mechanism of the
637
nitronium tetrafluoroborate nitration of alkylbenzenes in tetramethylene sulfone solution.
638
Remarks on certain aspects of electrophilic aromatic substitution2. J. Am. Chem. Soc. 1961, 83
639
(22), 4571–4580.
640
(43) Coombes, R. G.; Moodie, R. B.; Schofield, K. Electrophilic aromatic substitution. Part I.
641
The nitration of some reactive aromatic compounds in concentrated sulphuric and perchloric
642
acids J. Chem. Soc. B 1968, 800-804.
643
(44) Kenner, J. Oxidation and Reduction in Chemistry. Nature 1945, 156, 369-370.
644
(45) Weiss, J. Simple electron transfer processes in systems of conjugated double bonds. T.
645
Faraday Soc. 1946, 42, 116-121.
646
(46) Kochi, J. K. Inner-sphere electron-transfer in organic-chemistry – relevance to electrophilic
647
aromatic nitration. Acc. Chem. Res 1992, 25 (1), 39-47.
648
(47) Peluso, A.; DelRe, G. On the occurrence of an electron-transfer step in aromatic nitration.
649
J. Phys. Chem. 1996, 100 (13), 5303-5309.
650
(48) Albunia, A. R.; Borrelli, R.; Peluso, A. The occurrence of electron transfer in aromatic
651
nitration: dynamical aspects. Theor. Chem. Acc. 2000, 104 (3-4), 218-222.
652
(49) Challis, B. C.; Lawson, A. J.; Higgins, R. J. Chemistry of nitroso-compounds .3.
653
Nitrosation of substituted benzenes in concentrated acids. J. Chem. Soc., Perkin Trans. 2 1972,
654
(12), 1831-1836.
655
(50) Skokov, S.; Wheeler, R. A. Oxidative aromatic substitutions: Hartree-Fock/density
656
functional and ab initio molecular orbital studies of benzene and toluene nitrosation. J. Phys.
657
Chem. A 1999, 103 (21), 4261-4269. 32 ACS Paragon Plus Environment
Page 33 of 34
Environmental Science & Technology
658
(51) Chen, L. T.; Xiao, H. M.; Xiao, J. J.; Gong, X. D. DFT study on nitration mechanism of
659
benzene with nitronium ion. J. Phys. Chem. A 2003, 107 (51), 11440-11444.
660
(52) Frka, S.; Šala, M.; Kroflič, A.; Huš, M.; Čusak, A.; Grgić, I. Quantum Chemical
661
Calculations Resolved Identification of Methylnitrocatechols in Atmospheric Aerosols.
662
Environ. Sci. Technol. 2016, 50 (11), 5526-5535.
663
(53) Olah, G. A. Mechanism of eletrophilic aromatic substitution. Abstr. Pap. Am. Chem. Soc.
664
1970, 7 (4), 240-248.
665
(54) Hemberger, P.; da Silva, G.; Trevitt, A. J.; Gerber, T.; Bodi, A. Are the three
666
hydroxyphenyl radical isomers created equal? - The role of the phenoxy radical. Phys. Chem.
667
Chem. Phys. 2015, 17 (44), 30076-30083.
668
(55) Gowenlock, B. G.; Pfab, J.; Young, V. M. Kinetics of the oxidation of aromatic C-nitroso
669
compounds by nitrogen dioxide. J. Chem. Soc., Perkin Trans. 2 1997, (9), 1793-1798.
670
(56) Park, J. Y.; Lee, Y. N. Solubility and decomposition kinetics of nitrous acid in aqueous
671
solution. J. Phys. Chem. 1988, 92 (22), 6294-6302.
672 673
For Table of Contents only.
674
33 ACS Paragon Plus Environment
Environmental Science & Technology Page 34 of 34
ACS Paragon Plus Environment